2013 review tech qd-leds klimov

Upload: jpmartinep4318

Post on 03-Jun-2018

216 views

Category:

Documents


0 download

TRANSCRIPT

  • 8/12/2019 2013 REVIEW Tech QD-LEDs Klimov

    1/10

    721 2013 Materials Research Society MRS BULLETIN VOLUME 38SEPTEMBER 2013 www.mrs.org/bulletin

    Introduction: Traditional and emerging LED

    technologiesBeginning with the ancient myth of Prometheus, who stole fire

    from Mount Olympus, to third millennium technologies, the

    ability to master artificial light incarnates the very concept of

    progress and modernity. Throughout history, the impact of

    artificial lighting on human civilization has been so profound

    that it has arguably redefined our concepts of time and space,

    extending our productive time and allowing humankind to

    explore even the darkest corners of our world.

    Our society relies so heavily on lighting that it dedicates a

    fifth of global electricity to it, which amounts to over 2,650 TWh

    of energy. In 2010, the environmental cost associated with

    the production of this amount of energy was equal to abouttwo billion tons of carbon dioxide emitted in the atmosphere.1

    Unfortunately, most of this precious energy is wasted due to

    the inefficiency of incandescent bulbs that convert only 10%

    of electricity into visible light, while 90% is dissipated as heat.

    Fluorescent lamps are less wasteful but still have a conversion

    efficiency of only 2030%. The economic advantage of over-

    coming these limitations, together with a growing awareness

    of the harmful effects of carbon emission on global climate,

    has driven the tremendous effort to develop more efficient andsustainable lighting technologies.

    Direct conversion of electricity into light using semicon-

    ductor-based light-emitting diodes (LEDs) is now universally

    accepted as the most promising approach to more efficient

    lighting. Solid-state lighting (SSL) sources such as LEDs

    demonstrate high brightness, long operational lifetime, and low

    energy consumption, far surpassing the performance of conven-

    tional lighting systems. The LED field is currently dominated by

    semiconductor quantum-well emitters (based, e.g., on InGaN/

    GaN) fabricated via epitaxial methods on crystalline sub-

    strates (e.g., sapphire).1 These structures are highly efficient

    and bright, but their cost-to-light-output ratio (also referredto as the cost of light) is at least one-hundred times larger

    than that of incandescent light bulbs. In addition, traditional

    semiconductor LEDs are affected by structural defects at the

    substrate/semiconductor interface due to lattice mismatch and

    heating during operation, which limits the device area to only

    12 mm2. Furthermore, harsh conditions required for the fab-

    rication of these structures and difficulties in post-processing

    Spectroscopic insights into theperformance of quantum dotlight-emitting diodesWan Ki Bae, Sergio Brovelli, and Victor I. Klimov

    Lighting consumes almost one-fth of all electricity generated today. In principle, with more

    efcient light sources replacing incandescent lamps, this demand can be reduced at least

    twofold. A dramatic improvement in lighting efciency is possible by replacing traditional

    incandescent bulbs with light-emitting diodes (LEDs) in which current is directly converted

    into photons via the process of electroluminescence. The focus of this article is on the emerging

    technology of LEDs that use solution-processed semiconductor quantum dots (QDs) as light

    emitters. QDs are nano-sized semiconductor particles whose emission color can be tuned by

    simply changing their dimensions. They feature near-unity emission quantum yields and narrow

    emission bands, which result in excellent color purity. Here, we review spectroscopic studies

    of QDs that address the problem of nonradiative carrier losses in QD-LEDs and approaches

    for its mitigation via the appropriate design of QD emitters. An important conclusion of our

    studies is that the realization of high-performance LEDs might require a new generation of

    QDs that in addition to being efcient single-exciton emitters would also show high emission

    efciency in the multicarrier regime.

    Wan Ki Bae, Chemistry Division, Los Alamos National Laboratory; [email protected] Brovelli, Department of Materials Science, Universit degli Studi di Milano-Bicocca; [email protected] I. Klimov, Chemistry Division, Los Alamos National Laboratory; [email protected]

    DOI: 10.1557/mrs.2013.182

  • 8/12/2019 2013 REVIEW Tech QD-LEDs Klimov

    2/10

    SPECTROSCOPIC INSIGHTS INTO THE PERFORMANCE OF QUANTUM DOT LIGHT-EMITTING DIODES

    722 MRS BULLETIN VOLUME 38SEPTEMBER 2013 www.mrs.org/bulletin

    complicate their use in lightweight, flexible lighting devices and

    displays. These drawbacks of traditional LEDs have motivated

    the search for new technologies, such as LEDs based on

    organic semiconductors (OLEDs).

    OLEDs offer greatly reduced fabrication cost compared to

    inorganic LEDs and are easily amendable to low-temperature,

    large-area processing, including fabrication on flexible sub-

    strates. Synthetic organic chemistry provides essentially an

    unlimited number of degrees of freedom for tailoring molecu-

    lar properties to achieve specific functionality, from selective

    charge transport to color-tunable light emission. The prospect

    of high-quality lighting sources based on inexpensive plastic

    materials has driven a tremendous amount of research in the

    area of OLEDs, which in turn has led to the realization of

    several OLED-based high-tech products such as flat screen

    televisions and mobile communication devices. Several industri-

    al giants such as Samsung, LG, Sony, and Panasonic are work-

    ing on the development of large-area white-emitting OLEDs

    for replacing incandescence light bulbs and fluorescent lightsources.

    In the most common embodiment, an OLED has a multilayer

    structure comprising an emissive layer (typically an organic

    matrix doped with molecular chromophores) sandwiched

    between organic hole and electron transport layers.2,3 The

    emissive chromophores are usually either conjugated fluo-

    rophores or phosphorescent metal-organic complexes.4,5

    Conjugated fluorophores (e.g., 4-(Dicyanomethylene)-2-methyl-

    6-[p-(dimethylamino)styryl]-4H-pyran, coumarins and perylene

    derivatives)3 have been a major subject of OLED research for

    over two decades.68 They feature the high emission yields

    and short luminescence lifetimes typical of singlet emitters,which can convert the energy of singlet excitons (electron

    hole pairs with total spin 0) to photons. However, their draw-

    back is that their maximum electroluminescence (EL) internal

    quantum efficiency (IQE; the ratio of the number of photons

    emitted within a device to the number of electrons passing

    through the device) is limited to 25%, as 75% of excitons gen-

    erated by charge injection are triplets, which are nonemissive

    spin 1 species. To date, external quantum efficiencies (EQEs;

    the ratio of the number of photons emitted from a device to the

    number of electrons passing through the device) of over 10%

    have been obtained with red9,10 and green11 conjugated fluo-

    rophores that also demonstrate impressive device lifetimes inexcess of 150,000 hours. Blue OLEDs are more difficult to

    realize, and so far the demonstrated EQEs from fluorescent

    emitters do not exceed 56%, and the longest lifetimes are

    about 10,000 hours.12

    The more recently introduced phosphorescent metal-organic

    complexes13 allow one, in principle, to reach 100% IQE by over-

    coming the limits imposed by branching between spin states of

    different multiplicities. Red and green OLEDs based on phos-

    phorescent iridium complexes showed EQEs above 20% with

    a turn-on voltage below 3 V.14 More recently, green-emitting

    OLEDs with an optimized light out-coupling design demonstrat-

    ed an EQE of 63%.

    15

    Blue phosphorescent OLEDs also exhibit

    performance competing with that of the best devices based on

    blue fluorescent molecules, exhibiting EQEs as high as 14%.16

    Despite impressive advances in the OLED field, there are

    a few drawbacks of this technology that might prevent its

    widespread use in commercial products. One problem is poor

    cost-efficiency, which is at least partially due to the complex-

    ity of the necessary device architecture, which requires pro-

    cessing involving multiple thermal deposition steps. Another

    problem is their limited stability, particularly for deep-red and

    blue phosphorescent OLEDs. While being greatly improved

    over the years, it still does not meet the standards employed

    in high-end devices. Furthermore, the color purity of OLEDs,

    that is, the width of the emission band, is inferior to that of

    conventional semiconductor LEDs. Specifically, the emission

    full width at half maximum of OLEDs is typically greater than

    40 nm (Figure 1, dashed lines), while it is less than 30 nm for

    conventional LEDs.

    A promising class of emissive materials for low-cost yet

    efficient LEDs is chemically synthesized nanocrystal quantumdots (QDs). These luminescent nanomaterials feature size-

    controlled tunable emission wavelengths, combined with

    improvements in both color purity (Figure 1, solid lines) as

    well as stability and durability over organic molecules. At the

    same time, as with organic materials, colloidal QDs can be

    fabricated and processed via inexpensive solution-based tech-

    niques compatible with lightweight, flexible substrates.1719

    Similar to other semiconductor materials, colloidal QDs

    feature almost continuous above-band-edge absorption and a

    narrow emission spectrum at a near-band-edge energy. Distinct

    from bulk semiconductors, however, the optical spectra of QDs

    depend directly on their size. Specifically, their emission colorcan be continuously tuned from the infrared (IR) to ultraviolet

    Figure 1. High color quality of quantum dot (QD) emission.

    Comparison of typical electroluminescent spectra of blue, green,

    and red QD-LEDs (light-emitting diodes) (solid lines) to those of

    OLEDs emitting at similar wavelengths (dashed lines).

  • 8/12/2019 2013 REVIEW Tech QD-LEDs Klimov

    3/10

  • 8/12/2019 2013 REVIEW Tech QD-LEDs Klimov

    4/10

    SPECTROSCOPIC INSIGHTS INTO THE PERFORMANCE OF QUANTUM DOT LIGHT-EMITTING DIODES

    724 MRS BULLETIN VOLUME 38SEPTEMBER 2013 www.mrs.org/bulletin

    to dense QD films typically used as the active layers of LEDs.

    In this case, the role of surface recombination is amplified be-

    cause of energy transfer (ET), which allows for a given exciton

    to sample trap centers in multiple QDs within the ET distance

    (the distance for which the ET time is equal to the exci-

    ton recombination time in an isolated QD) (Figure 2b). In

    QDs, ET typically leads to a red shift of the PL peak, which

    is accompanied by acceleration of PL decay, especially pro-

    nounced on the blue side of the emission spectrum, which cor-

    responds to smaller dots in the sample.52,53 The emitting dipole

    of these dots readily couples to the strong absorption dipole

    of bigger dots, which leads to a migration of excitons to the

    largest QDs in the ensemble.

    For closely spaced CdSe QDs (e.g., in dense Langmuir

    Blodgett films54), the ET time can be as short as ca. 100 ps.

    It is slower in commonly used films made via drop cast-

    ing, dip coating, or spin coating; however, even in this case,

    the characteristic time of ET (typically a few nanoseconds52)

    is faster than the radiative lifetime (20 ns for CdSe QDs55).Therefore, it can compete with radiative recombination, lead-

    ing to reduced PL QYs compared to solution samples.

    To quantify this reduction, we consider a simplified situ-

    ation where the QD ensemble comprises only two types of

    dots: bright with PL QY equal to unity and dark with a

    zero QY. The corresponding QY (Q0) of a non-interacting QD

    ensemble is simply defined by the fraction of bright dots. If

    this sample is assembled in a film, the exciton generated in a

    bright QD can either recombine within it (radiatively under

    our assumptions) or migrate to a different dot with the total

    rate kET. Since the acceptor dot can be nonemissive (dark)

    with a probability (1Q0), each ET step leads to the reduction inthe PL QY. For the infinite number of these steps, the overall

    PL QY can be found from

    ( )

    ( ) ( ) }0 r r ET

    2

    0 ET r ET 0 ET r ET1 ,

    k k k

    k k k k k k

    = +

    + + + + +

    Q Q

    Q Q (1)

    which yields

    ( ( )0 r r ET 01 .k k k = + Q Q Q (2)

    If, for example, we consider a sample with Q0= 0.8 and typi-cal values of radiative (0.05 ns1) and ET (0.5 ns1) rates,

    we obtain Q = 0.27, three times lower than the original

    QY. This estimation shows that even highly luminescent

    solution samples may experience a significant reduction in

    effective emission efficiency upon processing into a film.

    This highlights the need for designing QD samples that fea-

    ture not only high emission efficiencies in solution form,

    but also suppressed ET when they are assembled into dense

    films.

    A recent study of LEDs based on CdSe/CdS QDs56 has

    demonstrated that an increase in the shell thickness (H) leads

    to a progressive increase in the LED EQE, which among other

    factors was attributed to suppression of exciton quenching by

    ET. Figure 2c displays the effect of increasing shell thickness

    on PL dynamics of QD film versus solution samples. As was

    pointed out earlier, the PL decay measured on the blue side

    of the emission band of the QD films represents a qualitative

    measure of the ET time. As evident from Figure 2c, this

    decay becomes progressively slower with increasing H, and

    in the thickest-shell sample (H= 6.4 nm), it is essentially the

    same as in the solution of non-interacting QDs. This suggests

    a nearly complete suppression of ET due to the thick shells

    acting as spacers separating the cores of the donor and the

    acceptor QDs involved in the ET process. The observed ET

    suppression is well correlated with the increase in the EL EQE

    of QD films incorporated into LEDs.56

    Hot carrier trapping/recombination and B-type QDblinkingMore recently, there has been increasing evidence from both

    single-dot57,58 and ensemble5961 measurements of the importanceof nonradiative recombination involving surface/interface trap-

    ping of hot, unrelaxed carriers (Figure 3a). In the absence

    of other nonradiative channels, the PL QY due to hot electron

    trapping/recombination can be described by Q= kcool/(kcool+ kesc),

    where kescis the rate of hot-electron escape from the QD, and

    kcoolis the rate of intraband relaxation to the emitting band-

    edge state. A distinct feature of this mechanism is that while

    leading to reduced PL intensity, it does notchange the emis-

    sion lifetime, which under the previous assumption is defined

    by radiative decay of a band-edge exciton. If both channels

    (band-edge and hot-carrier trapping) are active, the resulting

    QY can be expressed as Q= kcoolkr/[(kcool+ khot)(kr+ knr)].Hot-carrier trapping has been invoked to explain photocharg-

    ing of QDs under excitation with above-bandgap photons,5961

    and recently was used to explain one of the regimes (so-called

    B-type blinking) of PL intermittency (that is, switching of

    a QD between high- and low-emissivity states) observed

    in single-dot measurements.58 This regime is characterized by

    strong fluctuations in PL intensity that are not accompanied

    by any significant variations in PL lifetime. This is exactly

    the signature of a recombination channel due to hot-electron

    traps, which randomly turns on and off due to fluctuations in

    the traps occupancy.58

    Strong support for this model was provided by demonstra-tions of the control of B-type blinking by applied potential for

    QDs incorporated into an electrochemical cell.58 Specifically,

    it was observed that under positive bias, which corresponded

    to a lower Fermi level, the low-emissivity (OFF) periods were

    extended due to increasing time spent by the hot-electron trap

    in the unoccupied, active state (Figure 3b). On the other hand,

    under sufficiently negative potential, the B-type OFF periods

    were completely suppressed, as expected for the situation

    of a higher Fermi level, which leads to filling (that is, blocking)

    of hot-electron traps (Figure 3cd). This explanation of B-type

    blinking suppression due to filling of electron-accepting trap sites

    is consistent with previous observations of PL enhancement

  • 8/12/2019 2013 REVIEW Tech QD-LEDs Klimov

    5/10

    SPECTROSCOPIC INSIGHTS INTO THE PERFORMANCE OF QUANTUM DOT LIGHT-EMITTING DIODES

    725MRS BULLETINVOLUME 38SEPTEMBER 2013www.mrs.org/bulletin

    and blinking suppression due to electron-donating molecules62,63

    and n-doped substrates.64

    As with band-edge trapping, the B-type hot-electron trap-

    ping can be suppressed by growing a sufficiently thick shell.

    The analysis of blinking behaviors for CdSe/CdS QDs58 indi-

    cates that the occurrence of B-type blinking is reduced as the

    thickness of the CdS shell is increased, and it is completely

    eliminated in samples with shells comprising more than 15

    monolayers of CdS. These thick-shell QDs (often referred to

    as giant- or g-QDs) were either nonblinking or exhibited

    so-called A-type blinking due to charging/discharging of the

    QD core.58

    Interestingly, a closer inspection of lifetime tra-jectories indicated that some of the nonblinking dots, while

    showing stable emission intensity, were still characterized by

    significant fluctuations of the PL lifetime (lifetime blinking57)

    due to switching between neutral exciton and negative trion

    states; the charged state was highly emissive because of sup-

    pressed Auger recombination in this type of structure (as dis-

    cussed in the following sections).

    Nonradiative Auger recombination of chargedexcitonsAn important mechanism for nonradiative losses in QDs

    is associated with Auger recombination of multicarrier states,

    that is, states involving more than one exciton.

    The simplest of these states are singly charged

    excitons, or negative (Figure 4a) and positive

    (Figure 4b) trions (Xand X+, respectively).

    In addition to radiative or extrinsic defect-

    related processes, charged excitons can decay

    via nonradiative Auger recombination, a pro-

    cess in which the energy of an electronhole

    pair is converted not into a photon but instead

    is transferred to a third carrier. This process

    is inefficient in bulk forms of wide-gap semi-

    conductors, but due to relaxation of momen-

    tum conservation,65 it is extremely efficient

    in QDs, where it is characterized by a short

    (a few ps to hundreds of ps) time constant

    (A) that exhibits direct scaling with QD vol-

    ume (V), A V, sometimes referred to as

    V-scaling.66,67

    Charged states can be created within QDsunder light exposure (photocharging5961), due

    to direct electrical contact with a metal or

    semiconductor substrate, or can arise because

    of an imbalance between electron and hole

    injection currents in LED devices. There is

    ample evidence that charging has a significant

    detrimental impact on the performances of

    light-emitting systems relying on QDs.41,58,6782

    To quantify the effect of extra charges on QD

    emission, we assume that the radiative decay

    rate scales linearly with the number of available

    recombination pathways, that is, it is directlyproportional to the product of the electron (Ne) and the hole

    (Nh) occupancy of the QD:83,84

    ( )r e h e h 1r ; ,k N N N N k = (3)

    where k1ris the radiative decay rate of a neutral exciton (Ne= 1;

    Nh= 1). Based on this expression, the radiative rates of nega-

    tive and positive trions are k+1r= k1r= 2k1and corresponding

    PL QYs are

    ( ) ( )1 1 1 1A 1 1 1 1A2 2 and 2 2 . + += + = +k k k k k k Q Q (4)

    Since in standard core-only or thin-shell QDs the rate of Auger

    decay is several orders of magnitude higher than the radiative

    decay rate, the expression for QYs can be approximated by

    Q1 2k1/k1Aand Q+1 2k1/k+1A.

    The trion Auger lifetimes can be inferred from spectroscopic

    measurements of a neutral biexciton (a state comprising two

    electrons and two holes) decay rate (k2A), which is related to

    the Xand X+decay rates as k2A= 2(k1A+ k+1A). In core-only

    QDs that feature mirror symmetric conduction and valence

    bands, k1A= k+1A= k2A/4. Such a situation is realized, for

    example, in PbSe QDs for which the observed trion Auger

    lifetime is indeed approximately four times longer than that of

    Figure 3. Hot-carrier recombination centers and B-type quantum dot (QD) blinking.58

    (a) Carrier trapping by hot-electron (or hole) traps (D1; rate kesc) competes with intraband

    relaxation into the emitting band-edge state (ratekcool); this process is only active if the D1trap

    is unoccupied, that is, the Fermi level (FL) is below the trap level. (b) Hot electron trapping

    can lead to so-called B-type blinking when the uctuations in the photoluminescence (PL)

    intensity (black trace) are not accompanied by appreciable changes in the PL lifetime

    (red trace). The intensity and lifetime trajectories are shown for CdSe/CdS QDs (shell

    thickness is 3.6 nm) incorporated into an electrochemical cell. These specic data were

    collected for the positive potential of 0.8 V, which shifts the FL down in energy and

    increases the time spent by the D1trap in the unoccupied state. (c) The hot-electron

    trapping channel can be shut off by raising the FL, which leads to lling of the D1traps.

    (d) At sufciently high negative potential (1 V in the gure), the B-type blinking is completely

    suppressed.

  • 8/12/2019 2013 REVIEW Tech QD-LEDs Klimov

    6/10

    SPECTROSCOPIC INSIGHTS INTO THE PERFORMANCE OF QUANTUM DOT LIGHT-EMITTING DIODES

    726 MRS BULLETIN VOLUME 38SEPTEMBER 2013 www.mrs.org/bulletin

    a biexciton (2A). Recent measurements of chemically doped

    samples85 indicate that in PbSe QDs with Eg = 0.76 eV

    (R= 3.85 nm), the XAuger lifetime is 360 ps, very close tothe expected value of 416 ps predicted from the measured 2A

    of 104 ps.

    The symmetry between the Xand X+recombination path-

    ways can be distorted in QDs of IIVI semiconductors due to

    the significant difference between the conduction and valence

    bands. Specifically, a much higher density of valence-band

    states favors the positive trion Auger recombination channel

    (Figure 4b) for which the energy conservation requirement is

    easier to meet. Recent measurements86 have evaluated the

    biexciton and negative trion Auger lifetimes in thin-shell CdSe/

    ZnS QDs using time-resolved picosecond cathodolumines-

    cence. The results of these measurements are summarized inFigure 4c (+1Ais derived from the measured 1Aand 2A). One

    can see that in larger dots (Eg= 1.94 eV), +1Aand 1Aare close

    to each other (770 ps and 799 ps, respectively). However, as

    the dot size is decreased, the XAuger pathway gets progres-

    sively suppressed due to increasing separation between the

    conduction band states, so that in QDs with Eg= 2.34 eV, the

    Auger lifetime of the positive trion (126 ps) is ca. twice as fast

    as that of the negative trion (254 ps).

    The asymmetry between the Xand X+recombination path-

    ways is further enhanced in coreshell structures that feature

    a significant difference in effective localization radii (defined

    by the spatial extent of electronic wave functions) of electrons

    (Re) and holes (Rh). One example is thick-shell

    CdSe/CdS g-QDs.70,75 In these dots, the electron

    is delocalized over the entire nanostructure,

    while the hole is confined to the smaller core,

    and hence, Re is much greater than Rh. This

    enhances the Auger decay rate of the positive

    versus negative trion due to increased prob-

    ability of hole-hole scattering events (involved

    in the X+decay channel; Figure 4b) compared

    to electron-electron scattering (involved in the

    Xdecay channel; Figure 4a). For example,

    based on a recent experimental study of core

    shell CdSe/CdS QDs,68 in samples with core

    radius r= 1.5 nm and H= 5.5 nm, 1A= 5 ns

    and 2A= 0.35 ns (inset of Figure 4d). These

    values yield +1A = 0.81 ns, which suggests

    a very significant asymmetry between the X

    and X+Auger decay channels with 1A 6+1A.

    On the basis of the previous analysis of trionAuger lifetimes, we can estimate their emis-

    sion efficiencies. Using the data obtained for

    standard thin-shell CdSe/ZnS QDs and assum-

    ing a 20 ns neutral-exciton radiative time con-

    stant,55 we estimate that for the sample with

    Eg= 2.34 eV (Figure 4c), the negative trion PL

    QY is only 2.5%, and this is even lower (1.3%)

    for positive trions. The trion QY is increased

    for larger samples (e.g., Q1 Q+1 7% for

    QDs with Eg= 1.94 eV); however, it is still below 10%.

    These estimations indicate that QD charging can greatly

    reduce emission efficiency and hence the EQE of LEDs. Aswas shown in previous work,72,73,82 the use of g-QDs with

    an ultra-thick CdS shell allows for significant suppression of

    Auger recombination, which translates into increased trion

    lifetimes and hence emission yields; this was cited as one of

    the reasons for improved performance of LEDs comprising

    CdSe/CdS g-QDs.56 Indeed, for the thick-shell CdSe/CdS

    sample in Figure 4d, we estimate that the negative trion PL

    QY is 20%, which is considerably higher than in standard

    dots. However, we would like to point out that because of a

    significant asymmetry between the Xand X+Auger pathways

    in these structures, the X+QY is still low (3%) and is com-

    parable to that in standard QDs. This observation highlightsthe fact that while charging is in general detrimental to LED

    performance, the presence of extra holes is especially harm-

    ful, as positive trion lifetimes in many types of IIVI QDs can

    be significantly shorter than the lifetimes of negative trions.

    Suppression of Auger recombination via interfaceengineeringThe previous analysis indicates that Auger recombination

    represents a serious impediment to the realization of high

    emission efficiency in situations where the formation of multi-

    carrier states is difficult to avoid, such as high-intensity (and/or

    high-photon energy) optical excitation or electrical pumping

    Figure 4. Auger recombination and the effect of interfacial alloying on the Auger decay

    rate. (a) Auger recombination of a negative trion can be described in terms of Coulombscattering between the two conduction band electrons when one of them undergoes an

    interband transition, while the other is excited within the conduction band. (b) Similarly,

    Auger recombination of a positive trion can be described as Coulomb scattering between

    the two valence-band holes. (c) Auger lifetimes of biexcitons (2A), and negative (1A)

    and positive (+1A) trions in standard thin-shell CdSe/ZnS quantum dots (QDs) with

    different bandgaps (Eg). (d) Early time Auger dynamics of mult icarrier states for reference

    coreshell (C/S) CdSe/CdS QDs with an abrupt interface and core/alloy/shell (C/A/S) CdSe/

    CdSexS1x/CdS QDs with an alloyed interface (xis ca. 0.5). Both samples have the same core

    radius (r= 1.5 nm) and the same total radius (R= 7 nm). The thickness of the interfacial

    alloy is 1.5 nm.68 (e) Schematic illustration of synthesis of the C/S and C/A/S QD samples.

  • 8/12/2019 2013 REVIEW Tech QD-LEDs Klimov

    7/10

    SPECTROSCOPIC INSIGHTS INTO THE PERFORMANCE OF QUANTUM DOT LIGHT-EMITTING DIODES

    727MRS BULLETINVOLUME 38SEPTEMBER 2013www.mrs.org/bulletin

    when even a slight imbalance in electron and hole currents

    can lead to QD charging with extra carriers. Therefore, the

    development of approaches for suppression of Auger decay

    could greatly benefit LEDs.

    As was mentioned earlier, one approach to reducing Auger

    decay rates involves overcoating CdSe cores with exception-

    ally thick CdS shells.72,73 Originally, these structures were

    expected to exhibit suppressed Auger recombination based on

    V-scaling arguments and the reduced electronhole overlap

    due to a considerable mismatch in effective radii of electron

    and hole localization (Re>> Rh). However, as was pointed out

    in Reference 72, these factors alone could not explain the obser-

    vation of very rapid lengthening of 2Awith increasing shell

    thickness, suggesting that other structural characteristics of

    the QD such as the properties of its interfaces might play an

    important role in Auger recombination.

    Indeed, calculations by Cragg and Efros71 as well as a

    follow-up theoretical study by Climente et al.87 have sug-

    gested that the shape of the confinement potential also has asignificant effect on the Auger decay rate. Specifically, these

    studies demonstrated that by smoothing the confinement

    potential (via, e.g., alloying the interfacial layer), one can

    reduce the overlap between the initial and the final state of the

    carrier excited during Auger recombination and thus achieve

    an orders-of-magnitude reduction in the rate of this process.

    Recently, the problem of the effect of inter-

    facial alloying on Auger recombination was

    directly addressed by studying carrier dynam-

    ics in coreshell CdSe/CdS QDs with a well-

    defined CdSexS1x alloy incorporated between

    the core and the shell (Figure 4de).68,88 Thesestudies were enabled by a novel synthesis

    method,68 allowing much faster growth of a

    thick CdS shell compared to a more traditional

    successive ionic layer adsorption and reaction

    (SILAR) method,70,75,89 which was essential for

    avoiding unintentional alloying at the CdSe/

    CdS interface. Using this method, it was pos-

    sible to synthesize structures with either a

    sharp coreshell boundary (C/S QDs) or with a

    CdSexS1x alloy layer incorporated between the

    core and shell (C/A/S QDs); see Figure 4e.68

    Side-by-side studies of spectroscopic prop-erties of these QDs68 indicated that the incor-

    poration of the interfacial alloy layer did not

    significantly influence either spectra or dynamics

    of single excitons; however, it had a profound

    effect on multicarrier dynamics and PL QY

    of multicarrier states. For example, a direct

    comparison of the C/S and C/A/S samples

    (alloy layer thickness L= 1.5 nm and x 0.5)

    with the same core radius (r= 1.5 nm) and the

    same total size (R= 7 nm) indicates that the

    Auger lifetimes are systematically longer for

    the C/A/S QDs (Figure 4d). While observed

    for both Xand X+states, lifetime lengthening is especially

    pronounced for positive trions (2.8 ns versus 0.81 ns for the

    C/A/S and the C/S samples, respectively), which results in

    considerable improvement of PL QY (12% versus 3% in

    the sample without the interfacial alloy). A stronger effect of

    interfacial alloying on X+dynamics is expected based on the

    fact that it primarily affects the shape of the hole confinement

    potential. As discussed in the next section, the use of alloyed

    dots in LEDs allows one to improve their EQE compared to

    coreshell structures with a sharp interface, and this improve-

    ment is especially pronounced for high injection currents when

    the EQE loss due to Auger decay is particularly significant.

    Effects of charging in LEDs and high-intensityroll-off of LED effi ciencyWhen QDs are incorporated into devices, they can acquire a

    net charge due to interactions with electrodes and/or charge

    transport layers. Recently, this effect has been studied81 using

    the so-called inverted LED architecture, where a QD activelayer is sandwiched between an inorganic metal-oxide elec-

    tron transport layer and an organic hole transport layer based

    on conjugated molecules (Figure 5a). This architecture has

    been previously applied in high-performance devices2729,81,90

    that feature low turn-on voltage (nearly the same as the

    QD bandgap), high external EQEs at practical brightness

    Figure 5. An inverted quantum dot light-emitting diode (QD-LED) with organic

    and inorganic charge transport layers and the inuence of QD charging on device

    performance.81 (a) Schematic illustration of the structure of an inverted QD-LED with

    inorganic and organic charge transport layers. Electrons are injected from the ZnO/

    ITO side, and holes are injected from the CBP/MoOx/Al s ide.28 (b) An energy band

    diagram of a device with CdSe/CdSexS1x/CdS QDs forming an active emitting layer.

    (c) Photoluminescence (PL) decay dynamics of QDs in solution, on glass, and within

    the device. Inset: PL lifetimes for QDs within the device versus applied voltage. (d)

    External quantum efciencies (EQEs) versus current density for two QD-LEDs: one

    employing CdSe/CdS QDs with an abrupt in terface (black diamonds) and the other,

    CdSe/CdSexS1x/CdS QDs (xis ca. 0.5), with an alloyed interface (red circles). Both samples

    have the same core radius (r= 1.5 nm) and the same total radius (R= 7 nm); the thickness

    of the CdSexS1x alloy layer is 1.5 nm.

  • 8/12/2019 2013 REVIEW Tech QD-LEDs Klimov

    8/10

    SPECTROSCOPIC INSIGHTS INTO THE PERFORMANCE OF QUANTUM DOT LIGHT-EMITTING DIODES

    728 MRS BULLETIN VOLUME 38SEPTEMBER 2013 www.mrs.org/bulletin

    (1001000 cd/m2), and good operational stability (half life-

    time 500 hr).28,29

    Figure 5b shows an energy diagram of a device utilizing

    alloyed CdSe/CdSexS1x/CdS QD emitters (same samples as

    in Figure 4de) and an n-type colloidal ZnO electron injec-

    tion layer assembled on an indium-tin-oxide (ITO) cathode

    together with a 4,4-bis(N-carbazolyl)-1,1-biphenyl (CBP) hole

    conducting layer separated from the Al anode by a thin MoOxspacer.81 Due to the energetic proximity of the QD conduction-

    band level to the Fermi level of the ITO cathode, electron

    injection from ITO into QDs occurs spontaneously at zero bias,

    as indicated by a direct comparison of PL dynamics for dots

    assembled on the passive glass substrate versus the dots on

    the ITO/ZnO electrode (Figure 5c). In the former case, the PL

    dynamics are nearly the same as for isolated QDs in solution,

    while in the latter, the dynamics are significantly faster due to

    Auger recombination of the negative trion state arising from

    the combination of a photoexcited exciton and an injected elec-

    tron. Indeed, the measured PL lifetime of 8.5 ns is very close tothe Auger recombination time of a negative trion inferred from

    spectroscopic studies of QD solutions (inset of Figure 4d).

    The fact that the observed changes in carrier decay are due

    to excess electrons is confirmed by the observed changes in

    the PL lifetime as a function of applied bias (inset of Figure 5c).

    Under reverse bias, which makes electron injection less

    favorable, the PL decay becomes slower as expected for the

    situation in which a QD fluctuating between the charged and

    the neutral state spends an increasing amount of time in the

    neutral state.58 On the other hand, the PL dynamics get faster

    under direct bias due to increasing contribution from doubly

    negatively charged excitons.58

    A distinct asymmetry in the response to negative versus

    positive bias also indicates that the observed changes in PL

    dynamics are not significantly altered by effects that are in-

    dependent of the electric-field direction, such as field-induced

    spatial separation between a photo-generated electron and a

    hole with a QD.90 These observations also lead

    to the important conclusion that a likely reason

    for the efficiency roll-off (also known as effi-

    ciency droop) at large currents is the increasing

    degree of dot charging with excess electrons,

    which leads to increasing nonradiative carrier

    losses due to Auger recombination. Indeed, ifwe compare LEDs fabricated using coreshell

    CdSe/CdS QDs to those made using alloyed

    CdSe/CdSexS1x/CdS dots, which exhibit sup-

    pressed Auger recombination (Figure 4de),

    we observe that the onset of the EQE roll-off

    is considerably higher in alloyed QDs than

    in structures with a sharp coreshell interface

    (Figure 5d).

    Summary and outlookQD-LEDs have advanced tremendously over

    the past two decades, and progress has been

    especially fast recently due to significant improvements in the

    quality of QD materials as well as advances in the architec-

    ture of the devices. As illustrated in Figure 6, over the years,

    the QD-LED architecture has evolved from proof-of principle

    polymer-QD bilayer structures to modern devices employing

    direct charge injection from finely tuned electron and hole

    transport layers (see, e.g., Figure 5a). Recent refinements of

    this latter architecture have led to the demonstration of devices

    with a peak EQE of 18%,29 which is near the theoretical maxi-

    mum of 20%. At the same time, parallel engineering efforts

    have demonstrated the feasibility of full-color active-matrix-

    driven QD displays with patterned red, green, and blue QDs

    active layers.27

    As in the field of OLEDs, an important outstanding challenge

    in the field of QD-LEDs is the stability of device operation,

    especially for high injection currents. For example, the record

    efficiency devices of Reference 29 were able to maintain the

    highest EQE for only less than one hour. The reasons for

    the instability likely relate to both chemical (photochemical)degradation of the QDs as well as photophysical effects.

    As discussed in the previous sections, growing evidence

    suggests a detrimental effect of charging on LED perfor-

    mance. The presence of excess charges in the QDs might lead

    to both reversible degradation of the LED efficiency due to

    Auger recombination as well as irreversible changes due to

    activation of chemical reactions at the QD surface and/or in

    the ligand shell.41 As discussed earlier, charging is also a likely

    reason for the significant efficiency roll-off observed for high

    driving currents.28,29,81

    These considerations imply that realization of high-

    performance LEDs requires QD materials that exhibit highemission efficiencies not only in the single-exciton but also

    multi-carrier regimes. Two examples of such structures dis-

    cussed in this article are thick-shell CdSe/CdS g-QDs70,73,75

    and recently demonstrated CdSe/CdSexS1x/CdS QDs, with

    an intermediate alloyed layer introduced for suppression of

    Figure 6. History of quantum dot light-emitting diodes (QD-LEDs). The rst reported

    QD-LEDs were based on a QD/polymer bilayer structure.23After this demonstration of

    proof-of-principle devices, the performance of QD-LEDs has been persistently improved

    by employing various types of multilayer architectures with carefully engineered charge

    transport layers (CTLs). The most notable examples are QD-LEDs with QD active layers

    sandwiched between all-organic,92 all-inorganic,93 and hybrid organic/inorganic CTLs.94

    In addition to rapid advances in the efciency and durability of individual QD-LEDs, recent

    technological developments have yielded active matrix driven full-color QD displays with

    patterned red, green, and blue QDs active layers.27

  • 8/12/2019 2013 REVIEW Tech QD-LEDs Klimov

    9/10

    SPECTROSCOPIC INSIGHTS INTO THE PERFORMANCE OF QUANTUM DOT LIGHT-EMITTING DIODES

    729MRS BULLETINVOLUME 38SEPTEMBER 2013www.mrs.org/bulletin

    Figure 7. Operational principle and schematics of an InGaN/GaN light-emitting diode

    (LED)-backlit display with quantum dot (QD) color converters. (a) In a backlit display,

    blue emission from an InGaN/GaN quantum well (QW) LED is partially absorbed by

    green- and red-emitting QD down-converters; the mixture of three emitted colors results

    in trichromatic white emission. (b) A schematic diagram of a liquid-crystal display utilizing

    trichromatic QD-based backlight.

    Auger recombination.68 Other reported structures that dem-

    onstrated suppressed Auger decay include CdxZn1xSe/ZnSe78

    and CdTe/CdSe.91

    The detrimental effects of charging can also be reduced via

    a careful choice/design of charge transport layers for achiev-

    ing balanced electron and hole injection currents. Electrode

    engineering can also help lower the operational voltage, and,

    as a result, the strength of the electric field within the QD

    layer. This might also help mediate the problem of efficiency

    roll-off, as some studies suggest that in certain cases, this roll-

    off can be attributed to the reduction of the QD emission rate

    due to field-induced spatial separation between electrons and

    holes.90

    While the performance of QD-LEDs is quickly approaching

    industry standards for lighting and displays, the first commercialapplications of QDs will likely utilize them as down-converters.

    In such devices (e.g., backlit units), QDs are not excited by

    passing current, but instead are excited optically by absorb-

    ing blue radiation from InGaN/GaN quantum well LEDs

    before reemitting green or red light (Figure 7ab).3034 Recently,

    Sony Corporation has announced that it plans to incorporate

    QD-based down-conversion white LEDs as a backlight unit

    into their LCD television sets. Similar directions are being

    pursued by other big companies such as Samsung, LG, etc.

    Given these recent advances, it is reasonable to expect that the

    very exciting moment when the first QD product appears on

    the market is just around the corner.

    AcknowledgmentsThis work was supported by the Chemical Sciences, Biosciences,

    and Geosciences Division of Office of Science, Office of

    Basic Energy Sciences, US Department of Energy.

    References1. E.F. Schubert, Light-Emitting Diodes(Cambridge University Press, Cambridge,UK, 2006).2. S. Schols, Device Architecture and Materials for Organic Light-EmittingDevices: Targeting High Current Densities and Control of the Triplet Concentra-tion(Springer, London, UK, 2011).3. F. So, Organic Electronics: Materials, Processing, Devices and Applications(CRC Press, Boca Raton, FL, 2009).

    4. A.B. Tamayo, B.D. Alleyne, P.I. Djurovich, S. Lamansky,I. Tsyba, N.N. Ho, R. Bau, M.E. Thompson, J. Am. Chem.Soc.125, 7377 (2003).5. S. Lamansky, P. Djurovich, D. Murphy, F. Abdel-Razzaq, H.-E. Lee, C. Adachi, P.E. Burrows, S.R. Forrest,M.E. Thompson, J. Am. Chem. Soc.123, 4304 (2001).6. S.R. Forrest, M.L. Kaplan, P.H. Schmidt, J. Appl. Phys.56, 543 (1984).

    7. J.H. Burroughes, D.D.C. Bradley, A.R. Brown, R.N. Marks,K. Mackay, R.H. Friend, P.L. Burns, A.B. Holmes, Nature347,539 (1990).8. G. Gustafsson, Y. Cao, G.M. Treacy, F. Klavetter, N. Colaneri,A.J. Heeger, Nature357, 477 (1992).9. W.J. Begley, T.K.S.I. Hatwar, SID Intl. Symp. Dig. Tech.Papers37, 942 (2006).10. H. Kuma, J. Jinde, M. Kawamura, SID Intl. Symp. Dig.Tech. Papers38, 1504 (2007).11. M. Ricks, J.R. Vargas, K.P. Klubek, SID Intl. Symp. Dig.Tech. Papers38, 830 (2007).12. D. Kondakov, SID Intl. Symp. Dig. Tech. Papers39, 617(2008).13. M.A. Baldo, D.F. OBrien, Y. You, A. Shoustikov, S. Sibley,M.E. Thompson, S.R. Forrest, Nature395, 151 (1998).

    14. H. Sasabe, Y. Li, S. Su, T. Takeda, J. Kido, Jpn. J. Appl. Phys.46, 10 (2007).15. Z.B. Wang, M.G. Helander, J. Qiu, D.P. Puzzo, M.T. Greiner, Z.M. Hudson,

    S. Wang, Z.W. Liu, Z.H. Lu, Nat. Photonics5, 753 (2011).16. S.J. Yeh, M.F. Wu, C.T. Chen, Y.H. Song, Y. Chi, M.H. Ho, S.F. Hsu, C.H. Chen,Adv. Mater.17, 285 (2005).17. D.V. Talapin, J.S. Lee, M.V. Kovalenko, E.V. Shevchenko, Chem. Rev.110,389 (2010).18. A.P. Alivisatos, Endeavour21, 56 (1997).19. V.I. Klimov, Nanocrystal Quantum Dots(CRC Press, New York, 2010).20. L. Brus, J. Phys. Chem.90, 2555 (1986).21. A.P. Alivisatos, Science271, 933 (1996).22. C.B. Murray, C.R. Kagan, M.G. Bawendi, Annu. Rev. Mater. Sci.30, 545 (2000).23. V.L. Colvin, M.C. Schlamp, A.P. Alivisatos, Nature370, 354 (1994).24. V.I. Klimov, A.A. Mikhailovsky, S. Xu, A. Malko, J.A. Hollingsworth,C.A. Leatherdale, H.-J. Eisler, M.G. Bawendi, Science290, 314 (2000).25. O.E. Semonin, J.M. Luther, S. Choi, H.-Y. Chen, J. Gao, A.J. Nozik, M.C. Beard,Science334, 1530 (2011).26. G. Konstantatos, I. Howard, A. Fischer, S. Hoogland, J. Clifford, E. Klem,L. Levina, E.H. Sargent, Nature442, 180 (2006).

    27. T.-H. Kim, K.-S. Cho, E.K. Lee, S.J. Lee, J. Chae, J.W. Kim, D.H. Kim, J.-Y. Kwon,G. Amaratunga, S.Y. Lee, B.L. Choi, Y. Kuk, J.M. Kim, K. Kim, Nat. Photonics5,176 (2011).28. J. Kwak, W.K. Bae, D. Lee, I. Park, J. Lim, M. Park, H. Cho, H. Woo, D.Y. Yoon,K. Char, S. Lee, C. Lee, Nano Lett.12, 2362 (2012).29. B.S. Mashford, M. Stevenson, Z. Popovic, C. Hamilton, Z. Zhou, C. Breen,J. Steckel, V. Bulovic, M. Bawendi, S. Coe-Sullivan, P.T. Kazlas, Nat. Photonics7, 407 (2013).30. J. Lee, V.C. Sundar, J.R. Heine, M.G. Bawendi, K.F. Jensen, Adv. Mater.12,1102 (2000).31. E. Jang, S. Jun, H. Jang, J. Lim, B. Kim, Y. Kim, Adv. Mater.22, 3076 (2010).32. H. Woo, J. Lim, Y. Lee, J. Sung, H. Shin, J.M. Oh, M. Choi, H. Yoon, W.K. Bae,K. Char, J. Mater. Chem. C1, 1983 (2013).33. M. Achermann, M.A. Petruska, D.D. Koleske, M.H. Crawford, V.I. Klimov,Nano Lett.6, 1396 (2006).34. M. Achermann, M.A. Petruska, S. Kos, D.L. Smith, D.D. Koleske, V.I. Klimov,Nature429, 642 (2004).35. W.K. Bae, J. Kwak, J.W. Park, K. Char, C. Lee, S. Lee, Adv. Mater.21, 1690(2009).36. K.-S. Cho, E.K. Lee, W.-J. Joo, E. Jang, T.-H. Kim, S.J. Lee, S.-J. Kwon, J.Y. Han,B.-K. Kim, B.L. Choi, J.M. Kim, Nat. Photonics3, 341 (2009).37. T.-H. Kim, K.-S. Cho, E.K. Lee, S.J. Lee, J. Chae, J.W. Kim, D.H. Kim, J.-Y. Kwon,G. Amaratunga, S.Y. Lee, B.L. Choi, Y. Kuk, J.M. Kim, K. Kim, Nat. Photonics5,176 (2011).38. W.K. Bae, J. Kwak, J. Lim, D. Lee, M.K. Nam, K. Char, C. Lee, S. Lee, NanoLett.10, 2368 (2010).39. J.M. Caruge, J.E. Halpert, V. Wood, V. Bulovic, M.G. Bawendi, Nat. Photonics2, 247 (2008).40. V.I. Klimov, D.W. McBranch, C.A. Leatherdale, M.G. Bawendi, Phys. Rev. B60, 13740 (1999).41. M. Jones, S.S. Lo, G.D. Scholes, Proc. Natl. Acad. Sci. U.S.A.106, 3011 (2009).42. J.M. Luther, J.M. Pietryga, ACS Nano7, 1845 (2013).43. M.A. Hines, P. Guyot-Sionnest, J. Phys. Chem.100, 468 (1996).44. B.O. Dabbousi, J. Rodriguez-Viejo, F.V. Mikulec, J.R. Heine, H. Mattoussi,R. Ober, K.F. Jensen, M.G. Bawendi, J. Phys. Chem. B101, 9463 (1997).

  • 8/12/2019 2013 REVIEW Tech QD-LEDs Klimov

    10/10

    SPECTROSCOPIC INSIGHTS INTO THE PERFORMANCE OF QUANTUM DOT LIGHT-EMITTING DIODES

    730 MRS BULLETIN VOLUME 38SEPTEMBER 2013 www.mrs.org/bulletin

    45. X. Peng, M.C. Schlamp, A.V. Kadavanich, A.P. Alivisatos, J. Am. Chem. Soc.119, 7019 (1997).46. J.M. Pietryga, D.J. Werder, D.J. Williams, J.L. Casson, R.D. Schaller,V.I. Klimov, J.A. Hollingsworth, J. Am. Chem. Soc.130, 4879 (2008).47. W.K. Bae, J. Joo, L.A. Padilha, J. Won, D.C. Lee, Q. Lin, W.-K. Koh, H. Luo,V.I. Klimov, J.M. Pietryga, J. Am. Chem. Soc.134, 20160 (2012).48. D.V. Talapin, I. Mekis, S. Gtzinger, A. Kornowski, O. Benson, H. Weller,J. Phys. Chem. B108, 18826 (2004).

    49. R. Xie, U. Kolb, J. Li, T. Basch, A. Mews, J. Am. Chem. Soc.127, 7480 (2005).50. W.K. Bae, K. Char, H. Hur, S. Lee, Chem. Mater.20, 531 (2008).51. W.K. Bae, M.K. Nam, K. Char, S. Lee, Chem. Mater.20, 5307 (2008).52. S.A. Crooker, J.A. Hollingsworth, S. Tretiak, V.I. Klimov, Phys. Rev. Lett.89,186802 (2002).53. C.R. Kagan, C.B. Murray, M. Nirmal, M.G. Bawendi, Phys. Rev. Lett.76,1517 (1996).54. M. Achermann, M.A. Petruska, S.A. Crooker, V.I. Klimov, J. Phys. Chem.B107, 13782 (2003).55. S.A. Crooker, T. Barrick, J.A. Hollingsworth, V.I. Klimov, Appl. Phys. Lett.82, 2793 (2003).56. B.N. Pal, Y. Ghosh, S. Brovelli, R. Laocharoensuk, V.I. Klimov, J.A. Hollingsworth,H. Htoon, Nano Lett.12, 331 (2011).57. C. Galland, Y. Ghosh, A. Steinbrck, J.A. Hollingsworth, H. Htoon, V.I. Klimov,Nat. Commun.3, 908 (2012).58. C. Galland, Y. Ghosh, A. Steinbruck, M. Sykora, J.A. Hollingsworth, V.I. Klimov,H. Htoon, Nature479, 203 (2011).

    59. P.L. Redmond, L.E. Brus, J. Phys. Chem. C111, 14849 (2007).60. L.A. Padilha, I. Robel, D.C. Lee, P. Nagpal, J.M. Pietryga, V.I. Klimov, ACSNano5, 5045 (2011).61. J.A. McGuire, M. Sykora, I. Robel, L.A. Padilha, J. Joo, J.M. Pietryga,V.I. Klimov, ACS Nano4, 6087 (2010).62. V. Fomenko, D.J. Nesbitt, Nano Lett.8, 287 (2007).63. S. Hohng, T. Ha, J. Am. Chem. Soc.126, 1324 (2004).64. S. Jin, N. Song, T. Lian, ACS Nano4, 1545 (2010).65. J.M. Pietryga, K.K. Zhuravlev, M. Whitehead, V.I. Klimov, R.D. Schaller, Phys.Rev. Lett.101, 217401 (2008).66. V.I. Klimov, A.A. Mikhailovsky, D.W. McBranch, C.A. Leatherdale, M.G. Bawendi ,Science287, 1011 (2000).

    nDemand

    Climbing the Ladder of Density

    Functional Approximations

    Watch 2012 MRS Materials Theory Award Winner,John P. Perdew,Tulane University,give his talk FREEvia MRS OnDemand.

    www.mrs.org/f12-mta-video

    Custom Designed Systems

    for Sub-Kelvin SPMTrusted by Todays Leading Researchers toExplore Tomorrows Most Advanced Materials

    Temperatures Range: 0.009 K - 1 K

    Fully Bakeable Systems

    Solenoid Magnets up to 17 T

    Vector Magnets up to 9-4 T or 9-1-1T

    UHV Compatible: